Origins of the Loss of Concertedness in Pericyclic Reactions

Jun 9, 2000 - Sunaina Singh , Amit Aggarwal , Sebastian Thompson , João P. C. Tomé , Xianchun Zhu , Diana Samaroo , Mikki Vinodu , Ruomei Gao , and ...
0 downloads 16 Views 398KB Size
6078

J. Am. Chem. Soc. 2000, 122, 6078-6092

Origins of the Loss of Concertedness in Pericyclic Reactions: Theoretical Prediction and Direct Observation of Stepwise Mechanisms in [3 + 2] Thermal Cycloadditions Silvia Vivanco,† Begon˜ a Lecea,‡ Ana Arrieta,† Pilar Prieto,§ In˜ aki Morao,† Anthony Linden,| and Fernando P. Cossı´o*,† Contribution from the Kimika Fakultatea, Euskal Herriko Unibertsitatea, P.K. 1072, 20080, San Sebastia´ n-Donostia, Spain, Farmazi Fakultatea, Euskal Herriko Unibertsitatea, P.K. 450, 01080 Vitoria-Gasteiz, Spain, Facultad de Quı´mica, UniVersidad de Castilla-la Mancha, 13071 Ciudad Real, Spain, and Organisch-chemisches Institut der UniVersita¨ t Zu¨ rich, Winterthurstrasse 190, CH-8057, Zu¨ rich, Switzerland ReceiVed December 30, 1999

Abstract: Several [3 + 2] thermal cycloadditions between azomethine ylides and nitroalkenes have been studied both theoretically and experimentally. When the N-metalated 1,3-dipoles are used, the reaction is stepwise. The corresponding zwitterionic intermediates have been located computationally and observed by NMR monitoring. In the case of N-unsubstituted azomethine ylides, the reaction can be concerted or stepwise, depending upon the ability of the substituents to stabilize zwitterionic intermediates. A general model is proposed to explain the observed phenomena. This simple model can be extended to other thermal cycloadditions to predict the stepwise or concerted nature of their mechanisms without computing complete reaction coordinates.

Scheme 1a

Introduction The [3 + 2] cycloadditions constitute one of the most efficient and general methods for the synthesis of five-membered rings.1 Among the different versions of the reaction developed over the years, the [3 + 2] cycloadditions between azomethine ylides 1 and olefins 2 (Scheme 1) provide ready convergent access to pyrrolidine2 nuclei 3, whose importance in medicinal and synthetic organic chemistry is well-known. This reaction is particularly efficient when the azomethine ylide is stabilized by a carbonyl group. In this case, the corresponding [3 + 2] cycloaddition leads to highly substituted proline derivatives. There are at least three general methods for the generation of these azomethine ylides (Scheme 2). The first one consists of the conrotatory ring opening of conveniently substituted aziridines. This method was used by Huisgen3 to perform his seminal

a In this and other schemes, unless otherwise stated, the possible substituents at the different positions are not specified.

Scheme 2



Kimika Fakultatea. Farmazi Fakultatea. § Facultad de Quı´mica. | Organisch-chemisches Institut. (1) (a) Huisgen, R. Angew. Chem., Int. Ed. Engl. 1963, 2, 565. (b) 1,3Dipolar Cycloaddition Chemistry; Padwa, A., Ed; Wiley: New York, 1984; Vols. 1-2. (c) Cinquini, M.; Cozzi, F. In StereoselectiVe Synthesis; Helmchen, G.; Hoffmann, R. W.; Mulzer, J.; Schauman, E., Eds.; Georg Thieme: Stuttgart 1996; Vol. 5, pp 2953-2987. (d) Padwa, A. In ComprehensiVe Organic Synthesis; Trost, B. M.; Fleming, I., Eds.; Pergamon: Oxford, 1991; Vol. 4, pp 1069-1109. (e) Wade, P. A. In ComprehensiVe Organic Synthesis; Trost, B. M.; Fleming, I., Eds.; Pergamon: Oxford, 1991; Vol. 4, pp 1111-1168. (2) (a) Lown, J. W. In 1,3-Dipolar Cycloaddition Chemistry; Padwa, A., Ed.; Wiley: New York, 1984; Vol. 1, pp 653-732. (b) Grigg, R. Chem. Soc. ReV. 1987, 16, 89. (c) Tsuge, O.; Kanemasa, S. In AdVances in Heterocyclic Chemistry; Katritzky, A. R., Ed.; Academic Press: San Diego, 1989; Vol. 45, pp 232-349. (d) Vedejs, E. In AdVances in Cycloaddition; Curran, D. P., Ed.; Jai Press: Greenwich, 1988; Vol. 1, pp 33-51. (3) (a) Huisgen, R.; Scheer, W.; Mader, H. Angew. Chem., Int. Ed. Engl. 1969, 8, 602. (b) Huisgen, R.; Mader, H. J. Am. Chem. Soc. 1971, 93, 1777. (c) Hermann, H.; Huisgen, R.; Mader, H. J. Am. Chem. Soc. 1971, 93, 1779. (d) Huisgen, R.; Scheer, W.; Mader, H.; Brunn, E. Angew. Chem., Int. Ed. Engl. 1969, 8, 604. ‡

studies on the mechanism of the generation of 1,3-dipoles and their subsequent cycloaddition reactions with alkenes. The second method4 consists of the isomerization of an imine derivative of an R-aminoester containing an enolizable hydrogen to yield the azomethine ylide which can be trapped by reaction (4) (a) Grigg, R.; Kemp, J. J. Chem. Soc., Chem. Commun. 1978, 109. (b) Grigg, R.; Kemp, J. Tetrahedron Lett. 1980, 21, 1461. (c) Grigg, R. Bull. Soc. Chim. Belg. 1984, 93, 593.

10.1021/ja9945360 CCC: $19.00 © 2000 American Chemical Society Published on Web 06/09/2000

Loss of Concertedness in Pericyclic Reactions Scheme 3

J. Am. Chem. Soc., Vol. 122, No. 25, 2000 6079 no previous theoretical work has been reported for the thermal [3 + 2] cycloaddition between stabilized azomethine ylides and π-deficient alkenes. For the N-metalated version of the reaction only two papers, one from Kanemasa15 and one from our group16 have been reported. In these papers, SCF-MO studies using semiempirical Hamiltonians suggest stepwise mechanisms (Scheme 3, path b) for this version of the reaction. Within this context and in connection with our ongoing program on [3 + 2] cycloadditions,16,17 we report herein our study on the reaction between stabilized azomethine ylides and nitroalkenes to yield highly substituted proline derivatives. We have performed computational and experimental studies to exploit the synergy between both approaches. Finally, we have performed a theoretical study to elucidate the reasons underlying the transition from concerted to stepwise mechanisms. The model thus developed can be readily extended to any thermally allowed cycloaddition. Computational Methods

with the corresponding dipolarophile. Alternatively, the same substrate can be readily transformed into its N-metalated azomethine ylide with the appropriate salt in the presence of base.2c,5-7 This kind of reaction has proved to be remarkably efficient in its possible asymmetric versions.8 In addition, the reaction tolerates a wide range of substituents and is therefore well-suited for combinatorial synthesis.9 Given the importance of the reaction, the elucidation of its mechanism is crucial to understand the variables that determine its regio- and stereocontrol. After a vivid debate,10,11 it is now accepted that in general the reaction is concerted, according to the supra-supra mechanism proposed by Huisgen (Scheme 3, path a).12 Several years later, Huisgen himself13found a [3 + 2] cycloaddition whose mechanism was demonstrated to be stepwise, zwitterionic structures being the reaction intermediates. Finally, Schleyer et al.14 have recently shown that anionic [3 + 2] cycloadditions take place via stepwise mechanisms. However, (5) Kanemasa, S.; Tsuge, O. In AdVances in Cycloaddition Chemistry; Curran, D. P., Ed.; Jai Press: Greenwich, CT, 1993; Vol. 3, pp 99-159. (6) For recent examples, see: (a) Barr, D. A.; Dorrity, M. J.; Grigg, R.; Malone, J. F.; Montgomery, J.; Rajviroongit, S.; Stevenson, P. Tetrahedron Lett. 1990, 31, 6569. (b) Grigg, R.; Markandu, J.; Perrior, T.; Surendrakumar, S.; Warnock, W. J. Tetrahedron 1992, 48, 6929. (c) Grigg, R.; Montgomery, J.; Somasunderam, A. Tetrahedron 1992, 48, 10431. (7) (a) Kanemasa, S.; Yamamoto, H. Tetrahedron Lett. 1990, 31, 3633. (b) Kanemasa, S.; Tatsukawa, A.; Wada, E. J. Org. Chem. 1991, 56, 2875. (8) (a) Annunziata, R.; Cinquini, M.; Cozzi, F.; Raimondi, L.; Pilati, T. Tetrahedron: Asymmetry 1991, 2, 1329. (b) Patzel, M.; Galley, G.; Jones, P. G.; Chrapkowsky, A. Tetrahedron Lett. 1993, 34, 5707. (c) Galley, G.; Liebscher, J.; Patzel, M. J. Org. Chem. 1995, 60, 5005. (d) Rispens, M. T.; Keller, E.; DeLange, B.; Feringa, B. L. Tetrahedron: Asym 1994, 5, 607. (e) Waldmann, H.; Blaeser, E.; Jansen, M.; Letschert, H. P. Angew. Chem., Int. Ed. Engl. 1994, 33, 683. (f) Waldmann, H.; Blaeser, E.; Jansen, M.; Letschert, H. P. Chem. Eur. J. 1995, 1, 150. (g) Pyne, S. G.; Safaei, J.; Koller, F. Tetrahedron Lett. 1995, 36, 2511. (9) (a) Gong, Y. D.; Nadji, S.; Olmstead, M. M.; Kurth, M. J. J. Org. Chem. 1998, 63, 3081. (b) Hamper, B. C.; Dukesherer, D. R.; South, M. S. Tetrahedron Lett. 1996, 37, 3671. (c) Marx, M. A.; Grillot, A.-L.; Louer, C. T.; Beaver, K. A.; Barlett, P. A. J. Am. Chem. Soc. 1997, 119, 6153. (d) Murphy, M. M.; Schullek, J. R.; Gordon, E. M.; Gallop, M. A. J. Am. Chem. Soc. 1995, 117, 7029. (10) Houk, K. N.; Gonzalez, J.; Li, Y. Acc. Chem. Res. 1995, 28, 81. (11) (a) Huisgen, R. In 1,3-Dipolar Cycloaddition Chemistry; Padwa, A., Ed.; Wiley: New York, 1984; Vol. 1, pp 1-176. (b) Huisgen, R. J. Org. Chem. 1976, 41, 403. (c) Firestone, R. A.; J. Org. Chem. 1968, 33, 2285. (d) Firestone, R. A. Tetrahedron, 1977, 33, 3309. (12) (a)Huisgen, R. Angew. Chem., Int. Ed. Engl. 1963, 2, 633. (b) Huisgen, R. J. Org. Chem. 1968, 33, 2291. (13) (a) Huisgen, R.; Mloston, G.; Langhals, E. J. Am. Chem. Soc. 1986, 108, 6401. (b) Mloston, G.; Huisgen, R. Tetrahedron Lett. 1989, 30, 5373. (14) Neumann, F.; Lambert, C.; Schleyer, P. v. R. J. Am. Chem. Soc. 1998, 120, 3357.

All of the results presented in this work have been obtained by using the GAUSSIAN 9418 and GAUSSIAN 9819 series of programs, with the standard 3-21G(*), 6-31G*, and 6-31+G*20 basis sets. To include electron correlation at a reasonable computational cost, density functional theory (DFT)21 has been used. We have used the hybrid threeparameter functional developed by Becke,22 which is usually denoted as B3LYP. Zero-point vibrational energies (ZPVEs), when computed at the HF/3-21G and HF/6-31G* levels were scaled by 0.92 and 0.89, respectively.23 The ZPVEs computed using the B3LYP functional were not scaled. All transition structures have been fully optimized and characterized by harmonic analysis. For each located transition structure, only one imaginary frequency was found in the diagonalized Hessian matrix, and the corresponding vibration was found to be associated with nuclear motion along the reaction coordinate under study. In (15) Tatsukawa, A.; Kawatake, K.; Kanemasa, S.; Rudzinski, J. M. J. Chem. Soc., Perkin Trans. 2 1994, 2525. (16) Ayerbe, M.; Arrieta, A.; Cossı´o, F. P.; Linden, A. J. Org. Chem. 1998, 63, 1795. (17) (a) Morao, I.; Lecea, B.; Cossı´o, F. P. J. Org. Chem. 1997, 62, 7033. (b) Morao, I.; Cossı´o, F. P. J. Org. Chem. 1999, 64, 1868. (c) Cossı´o, F. P.; Morao, I.; Jiao, H.; Schleyer, P. v. R. J. Am. Chem. Soc. 1999, 121, 6737. (18) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Gill, P. M. W.; Johnson, B. G.; Robb, M. A.; Cheeseman J. R.; Keith, T.; Petersson, G. A.; Montgomery, J. A.; Raghavachari, K.; Al-Laham, M. A.; Zakrzewski, V. G.; Ortiz, J. V.; Foresman, J. B.; Peng, C. Y.; Ayala, P. Y.; Chen, W.; Wong, M. W.; Andres, J. L.; Replogle, E. S.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Binkley, J. S.; Defrees, D. J.; Baker, J.; Stewart, J. S.; HeadGordon, M.; Gonzalez, C.; Pople, J. A. Gaussian 94, revision B.2.; Gaussian, Inc.: Pittsburgh, PA, 1995. (19) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A., Jr.; Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K. N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.;. Mennucci, B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.;. Cui, Q.; Morokuma, K.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.;.Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Keith, T.; AlLaham, M. A.; Peng, C. Y.; Nanayakkara, A.; Gonzalez, C.; Challacombe, M.; Gill, P. M. W.;. Johnson, B.; Chen, W.; Wong, M. W.; Andres, J. L.; Gonzalez, C.; Head-Gordon, M.; Replogle, E. S.; Pople, J. A. Gaussian 98, revision A.5; Gaussian, Inc.: Pittsburgh, PA, 1998. (20) Hehre, W. J.; Radom, L.; Schleyer, P. v. R.; Pople, J. A. Ab Initio Molecular Orbital Theory; Wiley: New York, 1986; pp 76-87 and references therein. (21) Parr, R. G.; Yang, W. Density-Functional Theory of Atoms and Molecules; Oxford: New York, 1989. (22) (a) Kohn, W.; Becke, A. D.; Parr, R. G. J. Phys. Chem. 1996, 100, 12974. (b) Becke, A. D. J. Chem. Phys. 1993, 98, 5648. (c) Becke, A. D. Phys. ReV. A 1988, 38, 3098. (23) Pople, J. A.; Schlegel, H. B.; Krishnan, R.; De Frees, D. J.; Binkley, J. S.; Frisch, H.; Whiteside, R.; Hout, R. F., Jr.; Hehre, W. J. Int. J. Quantum Chem. Symp. 1981, 15, 269.

6080 J. Am. Chem. Soc., Vol. 122, No. 25, 2000 several significant cases intrinsic reaction coordinate (IRC)24 calculations were performed to connect unambiguously transition structures with reactants, cycloadducts and reaction intermediates. Bond orders and atomic charges were calculated with the natural bond orbital (NBO) method.25 Nonspecific solute-solvent interactions were simulated using either the Onsager-Kirkwood26,27 or the self-consistent isodensity polarization models (SCIPCM)28 as implemented by Wiberg.29 In several cases (vide infra) molecular mechanics computations were performed on several stationary points in order to determine the minimum energy conformations. These calculations were performed using the AMBER*30 force field as implemented in the MacroModel31 package. The different possible conformers were optimized and then Monte Carlo32 simulations were performed on 1000 structures. Atoms-in-molecules (AIM)33 calculations were performed with the AIMPAC34 program. Nucleus-independent chemical shifts (NICS),35 were calculated by means of the gauge-independent atomic orbitals36 (GIAO) method. The synchronicity parameter denoted in the text as Sy has been computed using the proposal of Perica`s and Moyano37 and according to the methodology previously reported by our group.38 For a perfectly synchronous reaction, Sy ) 1 and for a completely asynchronous reaction Sy ) 0.

ViVanco et al. Table 1. Relative Energiesa,b (kcal/mol) of the Conformations of Azomethine Ylides 1a,c

 ) 1.00

 ) 35.94

conformation 1a (X ) Li) 1c (X ) H) 1a (X ) Li) 1c (X ) H) A B C D

0.0 31.4 34.0 38.4

0.0 5.5 2.8 9.3

0.0 c 19.5 14.9

0.0 3.6 2.8 6.7

a Relative energies with respect to conformation A. b Values computed at the MP2/6-31+G*//HF/6-31+G* +∆ZPVE and the MP2(L1A1)/ 6-31+G*//HF(L1A1)/6-31+G* +∆ZPVE levels. c The starting geometries converged to conformation A upon optimization.

Results and Discussion Lithium Azomethines. We have studied first the structure and the properties of lithium azomethine 1a both in the gas phase and in acetonitrile solution ( ) 35.94).39 All of the possible conformations have been analyzed, and their relative energies are reported in Table 1. From our calculated energies it is concluded that these stabilized lithium azomethines exist exclusively in the conformation A, in which lithium interacts with both the nitrogen and oxygen atoms. Figure 1A shows the main geometric data of 1a in this conformation. An analysis of the energy densities H(rc) at the corresponding critical points40 reveals that the lithium-heteroatom interactions are electrostatic (24) Gonzalez, C.; Schlegel, H. B. J. Phys. Chem. 1990, 94, 5523. (25) (a) Reed, A. E.; Curtiss, L. A.; Weinhold, F. Chem. ReV. 1988, 88, 899. (b) Reed, A. E.; Weinstock, R. B.; Weinhold, F. J. Chem. Phys. 1985, 83, 735. (26) (a) Onsager, L. J. Am. Chem. Soc. 1936, 58, 1486. (b) Wong, M. W.; Wiberg, K. B.; Frisch, M. J. J. Am. Chem. Soc. 1992, 114, 523. (27) As one reviewer has indicated, this is an approximation of the solvation energy. Similarly, the  function is a very crude estimate of solvent polarity. See ref 46 for more detailed treatment of this problem. See also: Huisgen, R. Pure Appl. Chem. 1981, 53, 171 and references therein. (28) Tomasi, J.; Persico, M. Chem. ReV. 1994, 94, 2027. (29) (a) Wiberg, K. B.; Castejon, H.; Keith, T. A. J. Comput. Chem. 1996, 17, 185.(b) Wiberg, K. B.; Rablen, P. R.; Rush, D. J.; Keith, T. A. J. Am. Chem. Soc. 1995, 117, 4261. (30) Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K. M., Jr.; Ferguson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.; Kollman, P. A. J. Am. Chem. Soc. 1995, 117, 5179. (31) Macromodel, v5.0. Mohamadi, F.; Richards, N. G. J.; Guida, W. C.; Liskamp, R.; Lipton, M.; Caufield, C.; Chang, G.; Hendrickson, T.; Still, W. C. J. Comput. Chem. 1990, 11, 440. (32) (a) Chang, G.; Guida, W. C.; Still, W. C. J. Am. Chem. Soc. 1989, 111, 4379. (b) Saunders: M., Houk, K. N.; Wu, Y.-D.; Still, W. C.; Lipton, M.; Chang, G.; Guida, W. J. Am. Chem. Soc. 1990, 119, 1419. (33) Bader, R. F. W. Atoms in Molecules. A Quantum Theory; Clarendon Press: Oxford, 1990. (34) Cheeseman, J.; Keith, T.; Bader, R. F. W. AIMPAC v2.0; McMaster University: Hamilton, Ontario L8S 4M1, Canada. (35) Schleyer, P. v. R.; Maerker, C.; Dransfeld, A.; Jiao, H.; Hommes, N. J. R. V. E. J. Am. Chem. Soc. 1996, 118, 6317. (36) Wolinski, K.; Hilton, J. F.; Pulay, P. J. Am. Chem. Soc. 1990, 112, 8251. (37) Moyano, A.; Perica`s, M. A.; Valentı´, E. J. Org. Chem. 1989, 54, 573. (38) (a) Lecea, B.; Arrieta, A.; Roa, G.; Ugalde, J. M.; Cossı´o, F. P. J. Am. Chem. Soc. 1994, 116, 9613. (b) Lecea, B.; Arrieta, A.; Lopez, X.; Ugalde, J. M.; Cossı´o, F. P. J. Am. Chem. Soc. 1995, 117, 12314. (39) Reichardt, C. SolVents and SolVent Effects in Organic Chemistry VCH: Weinheim, 1990; pp 408-410. (40) Cremer, D.; Kraka, E. Croat. Chem. Acta 1989, 57, 1259.

Figure 1. B3LYP/6-31+G* (plain text) and B3LYP(L1A1)/6-31+G* ( ) 35.94, bold text) fully optimized geometry of the most stable conformation of lithium azomethine 1a. (A) Bond distances (Å) and NBO charges (au, boxed numbers). (B) Wiberg bond indices. In this and the following figures which incorporate ball-and-stick drawings, atoms are represented by increasing order of shadowing as follows: H, C, O, N.

Scheme 4

in nature.41 This result, together with the NBO analysis, leads to the valence-bond structure depicted in Figure 1B. According to these results, the structure of 1a is closer to that of an enolate than to that of an azomethine ylide. We have also studied the interaction between lithium azomethine 1a and ethylene 2a to yield the lithium amide 3aa (Scheme 4). This model reaction represents the simplest [3 + 2] cycloaddition between a lithium stabilized azomethine ylide and an alkene. The main geometric features of the stationary points found in the B3LYP/6-31+G* potential energy hypersurfaces, (41) (a) Lecea, B.; Arrieta, A.; Morao, I.; Cossı´o, F. P. Chem. Eur. J. 1997, 3, 20. (b) Lopez, X.; Ugalde, J. M.; Cossı´o, F. P. J. Am. Chem. Soc. 1996, 118, 2718.

Loss of Concertedness in Pericyclic Reactions

J. Am. Chem. Soc., Vol. 122, No. 25, 2000 6081 Scheme 5a

a

Only one enantiomer is drawn.

Figure 2. Main geometric features of the stationary points found in the 1a + 2a f 3aa reaction (see Scheme 4), computed in the gas phase and in acetonitrile solution. Bond distances are given in Å. Rp denotes the (3,+1) ring point of the electron density. See the caption of Figure 1. Table 2. Complexation Energiesa (∆Ec, kcal/mol), Activation Energiesb (∆Ea, kcal/mol), and Reaction Energiesc (∆Erxn, kcal/mol) of the 1a + 2a f 3aa Reactiond (see Scheme 4) method

∆Ec

∆Ea

∆Erxn

HF/6-31+G* MP2/6-31+G*e B3LYP/6-31+G* B3LYP(L1A1)/6-31+G*f

-6.6 -9.7 -7.3 -6.8

52.4 24.8 31.5 32.7

-3.9 -12.9 -0.4 -2.6

a Computed as ∆E ) E(6) - [E(1a) + E(2a)]. b Computed as c ∆Ea ) E(7) - E(6). c Computed as ∆Erxn ) E(3a) - [E(1a) + E(2a)]. d The zero-point energy corrections, conveniently scaled if necessary (see text) were computed at the geometry optimization level. e Singlepoint energy computed on the HF/6-31+G* optimized structures. f The energy differences in solution were computed in acetonitrile ( ) 35.94).

both in the gas phase and in simulated acetonitrile solution are depicted in Figure 2. The corresponding relative energies are reported in Table 2. The first step of this reaction is found to consist of the formation of an orientation complex resulting from the coordination of the lithium cation of 1a with the ethylene unit. From 6, the next stationary point is 7, that corresponds to a concerted supra-supra interaction between 1a and 2a. The AIM analysis of the electron density shows a (3,+1) ring point (Rp) in 7. In addition, the NICS computed at the ring point is found to be -20.44 ppm mol-1 at the SCF-GIAO/6-31+G*// B3LYP/6-31+G* level. This result indicates that this transition structure is in-plane aromatic.17 However, the computed synchronicity (Sy) for the 1a + 2a f 3aa transformation is calculated to be Sy ) 0.66 at the B3LYP/6-31+G* level. As can be seen in Figure 2, the large difference between the C4‚‚‚C5 and C3‚‚‚C2 bonds is responsible for this quite low synchronicity. In summary, the reaction between 1a and ethylene

Figure 3. Main geometric features of the stationary points found in the 1a + 2b f endo-3ab transformation (see Scheme 5). See the captions of Figures 1 and 2.

is predicted to be a concerted, although quite asynchronous, [3 + 2] cycloaddition, despite the considerable enolate-like character of 1a. This results in a quite large activation energy, as can be appreciated from Table 2. Therefore, it is expected that the substituent effect in this reaction could induce considerable changes in the reaction profile. Since the nitro group is one of the most versatile16,42 electronwithdrawing groups (EWG), we next computed the interaction between 1a and nitroethylene 2b (Scheme 5) to yield the cycloadducts 3ab. In this case, formation of two diastereomers is possible. We shall denote the endo structures as those in which both EWGs are in a cis relationship. If these groups are trans to each other, the corresponding structure shall be denoted exo (Scheme 5). (42) (a) Efremov, D. A.; Berestovitskaya, V. M.; Perekalin, V. V. Nitroalkenes; Conjugated Nitro Compounds; Wiley: Chichester, 1994. (b) Barrett, A. G. M.; Grabosky, G. G. Chem. ReV. 1986, 86, 751. For [3 + 2] cycloadditions between azomethine ylides and nitro-olefins see: Pak, C. S.; Nyerges, M. Synlett 1999, 1271 and references therein.

6082 J. Am. Chem. Soc., Vol. 122, No. 25, 2000

ViVanco et al.

Figure 4. Reaction profiles found in vacuo and in acetonitrile solution ( ) 35.94) for the 1a + 2b f endo-3ab transformation. Zero point vibrational energies have been included. The plain numbers correspond to the gas-phase results.

All of our attempts to locate concerted transition structures, as we did in the 1a + 2a f 3aa cycloaddition, were unfruitful. Instead, we found stepwise mechanisms in the B3LYP/6-31+G* potential energy surface, both in the gas phase and in acetonitrile solution. The stationary points which were located and characterized are shown in Figure 3, and the corresponding reaction profile is depicted in Figure 4. The first reaction intermediate is the result of the interaction between the lithium cation of 1a and one of the two oxygen atoms of the nitro group of 2b. 8 is ∼5 kcal/mol less stable in solution than in the gas phase (See Figure 4), but the complexation energy is still appreciable. We have also located a first transition structure endo-9 leading to the formation of the C2-C3 bond. This distance is slightly shorter than that found in 7. However, there is no appreciable interaction between C4 and C5 and the C2‚‚‚C3-C4 angle of attack is found to be ∼109° (See Figure 3). Therefore, this transition structure does not correspond to a concerted suprasupra mechanism but to a nucleophilic Michael-type addition in which 1a behaves as a lithium enolate. In addition, in endo-9 there is an electrostatic interaction between one of the oxygen atoms of the nitro group and the lithium atom. This additional stabilizing interaction is kept along the reaction coordinate thus leading to endo-3ab. The alternative exo stationary points are considerably higher in energy ( ∼10-20 kcal/mol along the reaction coordinate in the gas phase) since in those diastereo-

meric structures the Coulombic Li‚‚‚ONO interaction is not present (see Figures S1 and S2 of the Supporting Information). Therefore, a high stereocontrol can be expected for this reaction despite the stepwise nature of its mechanism. This is a remarkable prediction since stereochemical arguments have been crucial in the elucidation of concerted or stepwise mechanisms in other [3 + 2] cycloadditions.43 Another interesting point is that, according to our calculations, solvent effects accelerate this reaction step. Thus, in the 1a + 2b f endo-3ab transformation the activation barrier associated with the formation of the C2-C3 bond is calculated to be ∼1.33 kcal/mol lower than in vacuo (See Figure 4). The next point along the reaction coordinate is endo-10, in which the C2-C3 bond is formed. In this local minimum there is a considerable double bond character between C4 and the nitro group, with a NBO bond order value of 1.44 (B3LYP(L1A1)/6-31+G* level). The calculated charge for the nitro group is -0.96 e. Therefore, endo-10 corresponds to a zwitterionic intermediate in which the anionic part is a nitronate moiety and the positive part corresponds to an iminium cation. Since the charge separation is difficult in the gas phase, this (43) (a) Bihlmaier, W.; Geittner, J.; Huisgen, R.; Reissig, H.-V. Heterocycles 1978, 10, 147. (b) Van Auken, T. V.; Rinehart, K. L. J. Am. Chem. Soc. 1962, 84, 3736. (c) Houk, K. N.; Firestone, R. A.; Munchausen, L. L.; Mueller, P. H.; Arison, B. H.; Garcia, L. A. J. Am. Chem. Soc. 1985, 107, 7227.

Loss of Concertedness in Pericyclic Reactions Scheme 6a

J. Am. Chem. Soc., Vol. 122, No. 25, 2000 6083 Scheme 7a

a The usually isolated species correspond to the boxed structures. Ln represents solvent molecules or other ligands coordinated to the metal.

zwitterionic intermediate is highly stabilized in solution with respect to the gas phase (Figure 4), although the geometric features are quite similar in both media (Figure 3). The second transition structure endo-11 leads to the formation of the C4-C5 bond and its geometric features correspond to nitroaldol-like process,41a with a half-chair conformation for the Li-N1-C2-C5-O moiety (See Figure 3). Again, although the in vacuo and acetonitrile solution geometries of this saddle point are quite similar, the activation barrier is ∼7 kcal/mol larger in solution than in the gas phase, because of the large stabilization of endo-10 in solution. This result is due to the partial cancellation in the endo-11 structure of the charge separation present in the zwitterionic intermediate. Finally, we have characterized the N-lithiated cycloadduct endo-3ab, which lies 2.7 kcal/mol below endo-11 in acetonitrile solution (Figure 4). This reaction product is ∼8 kcal/mol less stable than endo-10 in solution, a result which explains why, in several cases, reaction products derived from intermediates of this kind have been reported.16,44 In this respect, it is noteworthy that in many experimental studies the N-metalated azomethines are generated using catalytic amounts of salts in the presence of bases such as tertiary amines.2b,c,8 Our results indicate that the reaction mechanism coupled with the catalytic cycle is that depicted in Scheme 6. Thus, the N-metalated azomethine ylide is generated in situ from an imine and then interacts with the alkene to yield the N-metalated cycloadduct via the corresponding zwitterionic intermediate. Then, the catalytic system is regenerated via proton transfer from the ammonium ion to the pyrrolidine ring. Since our calculations indicate that in the stepwise reaction between 1a and 2b the key intermediate endo-10 lies in a deep enough well, we thought that its evolution should be detectable by NMR monitoring, provided that the substituents present in the starting materials were appropriate. Therefore, we decided (44) (a) Tsuge, O.; Kanemasa, S.; Ohe, M.; Yorozu, K.; Takenaka, S.; Ueno, K. Chem. Lett. 1986, 1271. (b) Kanemasa, S.; Yoshioka, M.; Tsuge, O. Bull. Chem. Soc. Jpn 1989, 62, 869. (c) Werf, A.; Kellogg, R. M. Tetrahedron Lett. 1991, 32, 3727. (d) Kanemasa, S.; Kumegawa, E.; Wada, E.; Nomura, M. Bull. Chem. Soc. Jpn 1991, 64, 2990. (e) Chavan, S. P.; Venkatraman, M. S.; Shama, A. K.; Chittiboyina, A. G. Tetrahedron Lett. 1996, 37, 2857. (f) Rowley, M.; Leeson, P. D.; Williams, B. J.; Moore, K. W.; Baker, R. Tetrahedron 1992, 48, 3557.

a Reagents and conditions: i: LiClO (1 equiv), NEt (1.0 equiv), 4 3 CH3CN, 24 h, rt. ii: LiClO4 (1.0 equiv), NEt3 (1.0 equiv), CH3CN, 60 min, rt. iii: THF, HCl 2 N, 20 min, rt. iv: CF3CO2H, CH2Cl2, 5 h, rt. v: Ac2O (1.5 equiv), NEt3 (1.6 equiv), CH2Cl2, 48 h, rt. b In the chiral structures only one enantiomer is drawn.

to study the reaction between imine 1′b and nitroalkenes 2c,d which yields pyrrolidines 3bc and 3bd, respectively (Scheme 7). We chose this particular imine because, according to our computational studies, if the zwitterionic intermediates were formed, the two phenyl groups should stabilize the positive charge of the iminium fragment. Therefore, the second step, namely the nitroaldol-like process should be slower. We first performed the two cycloadditions, reported in Scheme 7, using lithium perchlorate as the lithium source and triethylamine as the base. The solvent was acetonitrile, the same one used in the calculations. After 24 h of reaction, the endo cycloadducts 3bc and 3bd were obtained in good yields. The endo stereochemistry was assigned on the basis of the values of the 3J2,3 coupling constants (4.0 Hz).16 In addition, the structure of endo3bc was confirmed by X-ray diffraction analysis45 (see Supporting Information). Therefore, the endo isomer was observed in these cases, as predicted by the calculations. We also monitored the reaction mixture of the 1′b + 2c f endo-3bc transformation at different times in order to observe the presence of the corresponding Michael intermediate. The results of this study are displayed in Figure 5A. As diagnostic signals we monitored the resonances of the methinic protons contiguous to the nitro group as shown in Figure 5B. The structure of intermediate 4bc was confirmed by isolation and derivatization to the stable and crystalline N-acyl derivative 5bc (See Scheme 7), whose structure was also determined by X-ray (45) Crystal data for endo-3bc: C25H24N2O5, Mr ) 432.47, orthorhombic, P212121, colorless prism (0.35 × 0.45 × 0.45 mm), a ) 10.797(2) Å, b ) 22.824(4) Å, c ) 8.768(3) Å, V ) 2160.7(8) Å3, Z ) 4, dcalc ) 1.329 g cm-3, µ ) 0.0932 mm-1, T ) 173(1) K, Mo KR radiation (λ ) 0.71069 Å), Rigaku AFC5R diffractometer, ω/2θ scans. Refinement on F gave R ) 0.0399, wR ) 0.0359, GOF ) 1.767 using 294 refined parameters and 2594 observed reflections with I > 2σ(I) from the 3333 collected (3219 unique) with 2θ < 55°. The structure was solved by direct methods and all non-hydrogen atoms were refined anisotropically.

6084 J. Am. Chem. Soc., Vol. 122, No. 25, 2000

ViVanco et al.

Figure 5. (A) Plot of the reaction mixture in the 1′b + 2c f endo-3bc reaction against the reaction time. (B) An example of the 300 MHz 1H NMR spectra of the reaction mixture showing the diagnostic signals of the different species.

analysis (see Supporting Information).46 Therefore, the curves of Figure 5A confirm that this reaction is stepwise as suggested by the previous computational study. To test if the first step of the reaction is irreversible, after 60 min of reaction, we added nitroalkene 2d, more electrophilic than 2c, to a reaction mixture of 1′b and 2c. After 24 h of reaction at room temperature, only cycloadduct endo-3bc was obtained, with no detectable amounts of endo-3bd. N-Protonated Azomethine Ylides. Since such compounds are usually generated in situ from the corresponding imines,2b,c,5 we have studied this previously reported transformation first. The main geometric features of imine 1′c and its azomethine ylide 1c are displayed in Figure 6. The transition structure 12 which connects both isomers via a 1,2-proton shift is also reported in the same Figure. The (46) Crystal data for 5bc: C14H18N2O6, Mr ) 310.30, monoclinic, P21/ c, colorless plate (0.14 × 0.32 × 0.45 mm), a ) 10.600(1) Å, b ) 15.514(2) Å, c ) 10.014(2) Å, V ) 1531.3(4) Å3, Z ) 4, dcalc ) 1.346 g cm-3, µ ) 0.106 mm-1, T ) 173(1) K, Mo KR radiation (λ ) 0.71069 Å), Rigaku AFC5R diffractometer, ω/2θ scans. Refinement on F gave R ) 0.0589, wR ) 0.0463, GOF ) 1.686 using 203 refined parameters and 1806 observed reflections with I > 2σ(I) from the 3838 collected (3506 unique) with 2θ < 55°. The structure was solved by direct methods and all nonhydrogen atoms were refined anisotropically.

corresponding activation energy is calculated to be 57.8 kcal/ mol in the gas phase (See Table 3), a quite large value. Inclusion of one p polarization function in the migrating hydrogen atom lowers the activation energy 2.7 kcal/mol. Another possibility for this transformation consists of the enolization of the imine to the corresponding enol imine 1′′c (Figure 6). In general, enolization is a complex process in which general acid and base catalysis47 as well as solvent48 and tunneling49 effects are important. We have computed the relative energies between 1′c and 1′′c in the gas phase and in several solvents. The results are included in Table 3. Although direct comparison with experimental values is not possible, the magnitude of the energy differences is in line with experimental values obtained on related compounds. For instance, it has been reported50 that 3-hydroxypyrrole is ∼1.2 kcal/mol less stable than 3-ketopyrrole in water at 298 K. (47) (a) Hart, H. Chem. ReV. 1979, 79, 515. (b) Bruice, P. Y.; Bruice, T. C. J. Am. Chem. Soc. 1976, 103, 1761. (48) Capon, B.; Rycroft, D. S.; Watson, T. W.; Zucco, C. J. Am. Chem. Soc. 1981, 103, 1761. (49) Bell, R. L.; Taveras, D. L.; Truong, T. N.; Simons, J. J. Comput. Chem. 1997, 63, 861. (50) Capon, B.; Guo, B.-Z.; Kwok, F. C.; Siddhanta, A. K.; Zucco, C. Acc. Chem. Res. 1988, 21, 135.

Loss of Concertedness in Pericyclic Reactions

J. Am. Chem. Soc., Vol. 122, No. 25, 2000 6085

Figure 6. Main geometric features of the stationary points found in the formation of 1,3-dipole 1c from imine 1c′. See the captions of Figures 1 and 2. Table 3. Keto-Enol Energiesa,b (∆EKE, kcal/mol), Enol-Azomethine Energiesa,c (∆EEY, kcal/mol), and Activation Energiesa,d (∆Ea1,2 and ∆Ea1,4, kcal/mol) Associated with the Formation of Azomethine 1c in Different Solvents solvent



∆EKE

∆EEY

∆Ea1,2

∆Ea1,4

gas phase

1.00

C6H5CH3

2.38

C6H5Br o-C6H4Cl2 C6H5CN

5.40 9.93 34.78

-0.3 (2.4)e -0.1 (2.7)e 0.1 0.3 0.4

5.5 (7.3)e 6.8 (8.4)e 8.0 8.8 9.1

55.2 (57.8)e 54.2 (56.8)e 53.4 53.0 52.6

12.3 (13.2)e 12.1 (12.6)e 11.8 11.6 11.5

a Computed at the B3LYP/6-31+G** + ∆ZPVE and the B3LYP(L1A1)/6-31+G** + ∆ZPVE levels in vacuo and in solution, respectively. b Computed as ∆EKE ) E(1′′c) - E(1′c). c Computed as ∆EEY ) E(13) - E(1c). d Computed as ∆Ea1,4 ) E(13) - E(1′′c) and ∆Ea1,2 ) E(12) - E(1′c), respectively. e Values calculated at the B3LYP/6-31+G* + ∆ZPVE and the B3LYP(L1A1)/6-31+G* + ∆ZPVE levels in vacuo and in solution, respectively.

We have located and characterized the transition structure 13 which connects 1′′c with 1c. As can be seen from Figure 6, both 1c and 1′′c correspond to the two minima of the intramolecular O‚‚‚H‚‚‚N bond, with the enol being favored (see Table 3). Therefore, our calculations suggest that the generation of N-unsubstituted azomethine ylides takes place via keto-enol tautomerization of the starting imines followed by isomerization

via transition structures related to 13. Given the lack of experimental data, it is difficult to assess which step is kinetically more relevant, although on the basis of experimental studies on related systems,47,48,50 it can be suggested that generation of 1c is the limiting step. The conformational equilibrium of azomethine ylide 1c is similar to that previously found for 1a, although the energy differences are smaller. Again, the conformation A is the most stable one, both in the gas phase and in solution (see Table 1), mainly because of the intramolecular hydrogen bond present in this conformer. The chief geometric features of 1c are also reported in Figure 6. Interestingly, the coefficients of the HOMO of 1c are those that might be expected for a 1,3-dipole, with a ratio between C1 and C3 of

Sc3 f3 3 ) - ) - ) 1.13 c1 f S 1

1

where c1 and c3 are the expansion coefficients of the 1 and 3 carbon atoms (Figure 6). Similarly, fi and Si are the nucleophilic Fukui functions and local softnesses of the same atoms, respectively. In the case of 1a, this ratio was found to be 1.35. This result and the NBO charges indicate that 1c is closer to a standard 1,3-dipole than 1a. Next, we studied the interaction between dipole 1c and ethylene 2a to yield pyrrolidine 3ca (Scheme 8). The stationary

6086 J. Am. Chem. Soc., Vol. 122, No. 25, 2000

ViVanco et al. Scheme 9a

a

In the chiral structures only one enantiomer is drawn.

Figure 7. Main geometric features of the stationary points found in the 1c + 2a f 3ca reaction (Scheme 8). See the captions of Figures 1 and 2.

Scheme 8

Table 4. Complexation Energiesa (∆Ec, kcal/mol), Activation Energiesb (∆Ea, kcal/mol), and Reaction Energiesc (∆Erxn, kcal/mol) of the 1c + 2a f 3ca Reactiond (see Scheme 8) method

∆Ec

∆Ea

∆Erxn

HF/6-31+G* MP2/6-31+G*e B3LYP/6-31+G* B3LYP(L1A1)/6-31+G*f

-1.9 -4.0 -1.7 +7.2

26.2 7.2 13.7 13.9

-50.2 -47.8 -33.1 -26.0

Computed as ∆Ec ) E(14) - [E(1c) + E(2a)]. b Computed as ∆Ea ) E(15) - E(14). c Computed as ∆Erxn ) E(3ca - [E(1c) + E(2a)]. d The zero-point energy corrections, conveniently scaled if necessary (see text) were computed at the geometry optimization level. e Single-point energy computed on the HF/6-31+G* optimized structures. f The energy differences in solution were computed in acetonitrile ( ) 35.94). a

points found in the B3LYP/6-31+G* energy hypersurfaces, both in the gas phase and in solution, are reported in Figure 7, and the corresponding energy differences are shown in Table 4. According to our results, the stability of the orientation complex 14 with respect to the reactants is quite low in the gas phase. When the complexation energy is computed in acetonitrile solution, inclusion of ZPVE results in a positive value (if the ZPVE correlation is not included, 14 lies 0.2 kcal/mol below

Figure 8. Chief geometric features of the stationary points found in the 1c + 2b f 3cb reaction (Scheme 9). See the captions of Figures 1 and 2.

the separate reactants 1c and 2a). Therefore, although the computed ∆Ea values reported in Table 4 are quite similar in the gas phase and in solution, if the activation energy is estimated from the separate reagents, the energy barrier is significantly larger in solution. The structure of 15 is depicted in Figure 7. This transition structure is also concerted, with a computed synchronicity of 0.79 in the gas phase. In addition, the computed NICS at the (3,+1) ring point is -21.30 ppm, which is an indicator of an aromatic transition structure. Then we studied computationally the interaction between azomethine ylide 1c in its A conformation and nitroethylene 2b. In this case, two diastereomers can be formed, endo-3cb and exo-3cb (Scheme 9). The stationary points located in the gas phase and in solution are shown in Figure 8, and the main

Loss of Concertedness in Pericyclic Reactions

J. Am. Chem. Soc., Vol. 122, No. 25, 2000 6087

Table 5. Complexation Energiesa,b (∆Ec, kcal/mol), Activation Energiesa,c (∆Ea, kcal/mol), Reaction Energiesa,d (∆Erxn, kcal/mol), Synchronicitiesa (SY), NICSe (ppm/mol), and TS-Dipole Momentsa (µTS, au) of 1c + 2b f 3cb Reaction 1c + 2b f endo-3cb

1c + 2b f exo-3cb

magnitude

 ) 1.00

 ) 35.94

 ) 1.00

 ) 35.94

∆Ec ∆Ea ∆Erxn Sy NICS µTS

-7.3 5.8 -35.1 0.78 -19.98 2.54

+6.6 14.4 -24.0 0.63 -18.44 3.40

-7.3 5.2 -35.4 0.76 -18.44 1.33

+6.6 17.3 -22.2 0.61 -15.86 2.12

a Computed at the B3LYP/6-31+G* + ∆ZPVE and the B3LYP(L1A1)/6-31+G* + ∆ZPVE levels in the gas phase and in solution, respectively. b Computed as ∆Ec ) E(16) - [E(1c) + E(2b)]. c Computed as ∆Ea ) E(17) - [E(1c) + E(2b)]. d Computed as ∆Erxn ) E(3cb) - [E(1c) + E(2b)]. e Computed at the GIAO-SCF/6-31+G*// B3LYP/6-31+G* and the GIAO-SCF/6-31+G*//B3LYP(L1A1)/631+G* levels in the gas phase and in solution, respectively.

Scheme 10a

a

In endo-3db only one enantiomer is drawn.

energetic and electronic features of the whole process are reported in Table 5. The shape of 16 resembles that found for 8 since the interaction between the azomethine ylide and the alkene takes place through one oxygen atom of the nitro group. However, the energy of complexation is again positive in solution when the ZPVE is taken into account (Table 5). Without this correction 16 lies 0.92 kcal/mol below the separate reactants at the B3LYP(L1A1)/6-31+G* in acetonitrile solution. Therefore, we can conclude that this complex is not chemically relevant under the usual experimental conditions. We have located two diastereomeric transition structures, exo17 and endo-17, that correspond to supra-supra concerted mechanisms. It is interesting to note that if solvent effects are considered, formation of endo-3cb is favored under both kinetic and thermodynamic control (see Table 5). In addition, endo-17 is more aromatic and more synchronous than its exo analogue. As can be seen from the geometric data reported in Figure 8, these transition structures are quite asynchronous, particularly in solution. Thus, the calculated synchronicities are found to be ∼0.6 for both process, the exo-pathway being slightly more asynchronous (See Table 5). Therefore, both transition structures correspond to almost “halfway” process in terms of synchronicity. This feature is readily appreciated if we examine the evolution of the C2‚‚‚C3 and C4‚‚‚C5 bond distances on passing from endo- and exo-17 to endo- and exo-3cb. In Figure S3 of the Supporting Information we have included the plot of these distances against the intrinsic reaction coordinate (IRC). From these data it can readily be seen that at IRC ≈ 1.2 au the C2C3 bond is already formed, whereas the C4‚‚‚C5 distance is still larger than 2 Å. Therefore, we conclude that this example is a limiting case between a concerted and a stepwise mechanism. Inclusion of adequate substituents at the 1,3-dipole could result in stepwise processes. To confirm this hypothesis, we have studied the reaction between nitroethylene 2b and azomethine ylide 1d to form compound endo-3db (Scheme 10). As we had previously found in the metalated version of the reaction, we expected that

inclusion of two phenyl groups in the starting azomethine could stabilize the corresponding zwitterionic intermediates. Given the size of the system, the search of the stationary points and the geometric optimizations were performed on the HF/3-21G and HF(L1A1)/3-21G energy hypersurfaces. The lowest energy conformations of several structures were located by Monte Carlo simulations using the AMBER* force field. The single point energies were then computed both in the gas phase and in acetonitrile solution at the B3LYP/6-31+G* and B3LYP(L1A1)/ 6-31+G* levels, respectively. The reaction profiles thus characterized are depicted in Figure 9. We have found that under the substituent effect the reaction turns out to be stepwise. The first transition structure, 18 is associated with the formation of the C2‚‚‚C3 bond, the corresponding distance being 2.272 Å in acetonitrile solution. The calculated C4-C5 distance is 3.253 Å in solution, which is an indicator of no appreciable bonding interaction between these atoms. We have also located a zwitterionic intermediate 19 in which the C2-C3 bond is completely formed and there is an intramolecular hydrogen bond between the O7 atom and the hydrogen atom bonded to N1. It is interesting to note that this intermediate was located only in solution, since in the gas phase the optimized structure has the hydrogen atom bonded to O7, i.e., the nitro group is in its aci form (Figure 9). The nitro tautomer of the latter structure, denoted as 22 in Figure 9 lies ∼25 kcal/mol below endo-19 in the gas phase. Another isomer of this zwitterionic intermediate is 20, which possesses an O7-C5 bond and is slightly more stable than endo19 in solution (see Figure 9). This structure resembles those experimentally found by Mayr et al. during their studies on [4 + 3] cycloadditions between 1,3-dipoles and dienes.51 Finally, we found the second transition structure endo-21, in which C4‚‚‚C5 is found to be 2.439 Å in acetonitrile solution. An IRC calculation in acetonitrile solution showed that this saddle point connects the cycloadduct endo-3db with 20. We performed several series of experiments to test the main predictions made by our computations on model compounds. First, we studied the variation of the reaction rate of the 1′b + 2c f 3bc process with the solvent. As in the metal-assisted version of the reaction, only the endo cycloadduct was obtained. We measured these reaction rates by monitoring the decay of the π f π* UV absorption of the starting nitroalkenes (Figure S4 of the Supporting Information). Since the measurement of the required absorbance was difficult at large reaction times, we used the Kezdy-Swinbourne method to determine the A∞ values.52 The plots used for the determination of A∞ and kobs in toluene solution are also shown in Figure S4. The observed kinetic constants are reported in Table 6 and show an acceleration of the reaction with increasing polarity of the solvent. The corresponding logarithmic values against the Onsager function ( - 1)/(2 + 1) are shown in Figure S5. These experimental findings are in line with the computed solvent dependence of the formation of azomethine ylide 1c from the enol derivative 1′′c via 13. In contrast, Huisgen et al.53 found that in solvents of increasing polarity [3 + 2] dipolar cycloadditions between alkenes and nitrones exhibit a moderate rate deceleration. (51) Mayr, H.; Baran, J.; Heigl, U. W. Gazz. Chim. Ital. 1991, 121, 373. (52) Espenson, J. H. Chemical Kinetics and Reaction Mechanisms; McGraw-Hill: New York, 1981; pp 24-27. (53) Geittner, J.; Huisgen, R.; Reissig, H.-V. Heterocycles 1978, 11, 109. (b) Huisgen, R.; Seidl, H.; Bru¨ning, I. Chem. Ber. 1969, 102, 1102. (c) Eckell, A.; George, M. V.; Huisgen, R.; Kende, A. S. Chem. Ber. 1977, 110, 578. See also: Bo¨hm, T.; Elender, K.; Riebel, P.; Troll, T.; Weber, A.; Sauer, J. Tetrahedron 1999, 55, 9515.

6088 J. Am. Chem. Soc., Vol. 122, No. 25, 2000

ViVanco et al.

Figure 9. Reaction profiles found in vacuo and in acetonitrile solution ( ) 35.94) for the 1d + 2b f 3db reaction (Scheme 10). See the captions of Figures 1 and 2. Table 6. Observed First-Order Rate Constants kobsa,b for the 1′b + 2c f 3bc Reaction in Different Solvents solvent

( - 1)/(2 + 1)c

104 kobs (s-1)

-ln kobs

C6H5CH3 C6H5Br o-C6H4Cl2 C6H5CN

0.24 0.37 0.43 0.47

0.69 ((0.03) 1.21 ((0.05) 1.77 ((0.12) 1.75 ((0.13)

9.58 9.02 8.64 8.65

Scheme 11a

a Computed by means of the Kezdy-Swinbourne method (see text). Kinetic measurements were carried out at 112 °C in a 0.02 M solution of the nitroalkene 2c and with an excess of 10 equiv of 1′b. c Onsager function where  stands for the dielectric constant of the solvent. b

Therefore, our results indicate that formation of the dipole is a relevant step in the thermal [3 + 2] cycloaddition between imines and π-deficient alkenes. In addition, we have detected by 1H NMR the formation of 4bc in the reaction between 1′b and 2c to form endo-3bc, as we did in the metalated version of this reaction. The structure of 4bc was unambiguosly determined by transformation to its acetamide 5bc (vide supra). Thus, after 20 and 46 h in refluxing toluene the endo-3bc:4bc ratios were 83:17 and 67:33, respectively. This is in agreement with our computational results on the stepwise nature of this reaction and on the relative energies of endo-3db and 22. In addition, these results indicate that the conversion from 19 to 22 (not included in our computational study) has a larger activation barrier than the 19 f 3db transformation. We have also studied the thermal reaction between imines 1′d-g and nitroalkenes 2c-g to yield the cycloadducts depicted in Scheme 11. In these cases, almost equimolar amounts of endo- and exo-cycloadducts were obtained. The structures of these cycloadducts were unequivocally assigned by comparison

a In the chiral compounds only one enantiomer is drawn. b Reagents and conditions: i: Toluene, 112 °C.

with structurally related compounds.16 In addition, we have not detected Michael cycloadducts in these reactions. This result suggests that these less substituted azomethine ylides react via mechanisms closer to that described in the study on the 1c + 2b f 2cb reaction (Scheme 9), the relative energies of the corresponding endo- and exo transition states being quite similar.

Loss of Concertedness in Pericyclic Reactions

J. Am. Chem. Soc., Vol. 122, No. 25, 2000 6089 Table 7. Intrinsic Reaction Coordinate Values (IRC au),a Internuclear Distances (Rij, Å),b Overlap Integrals (Sij, au),c Repulsive Energies (∆Erep, kcal/mol),d Michael Energies (∆ENu, kcal/mol),e and Huisgen-Woodward-Hoffmann Energies (∆Econ, kcal/mol)f for Selected Model [3 + 2] Reactions

Scheme 12

Origins of the Concerted and Stepwise Mechanisms. In the previous sections we have found that different mechanisms can operate in a similar transformation and that the stereochemical criterion is not always a useful indicator of the nature of the mechanism. In principle, there are two limiting mechanisms for the [3 + 2] cycloaddition between azomethine ylides and alkenes. In the concerted mechanism, there are bonding interactions between the C2 and C3 atoms, as well as between C4 and C5 (Scheme 12). From second-order perturbation theory,54 the following expression is obtained for the initial stages of the concerted process:

(cH2 cL3 β23 + cH5 cL4 β45)2

BfA ∆Econ ) ∆EAfB orb + ∆Eorb ) 2

HA - LB 2

(cL2 cH3

β23 + cL5 cH4 HB - LA

β45)

∆ENu ) ∆ECoul +

)

q2qi

∑ i∈B R

+2

HA - LB

R45

-3.978 -2.994 -1.897 -0.897

3.511 3.290 3.055 2.797

0.017 0.026 0.041 0.066

3.578 3.301 2.984 2.684

0.03 0.05 0.07 0.10

-0.05 -0.13 -0.38 -1.06

0.30 0.27 0.23 0.22

-3.932 -2.988 -1.894 -0.896

3.455 3.248 2.979 2.677

0.019 0.028 0.047 0.081

1c + 2b f endo-3cb 3.654 0.012 0.19 -1.52 3.457 0.018 0.45 -1.77 3.190 0.032 1.27 -2.03 2.941 0.051 3.60 -2.35

-0.11 -0.25 -0.72 -2.04

0.43 0.43 0.43 0.45

-3.960 -2.986 -1.894 -0.986

3.469 3.272 3.029 2.730

0.018 0.027 0.043 0.074

1a + 2b f endo-3ab 3.833 0.008 0.15 -1.30 3.636 0.013 0.34 -1.54 3.448 0.019 0.85 -1.83 3.314 0.025 2.38 -2.25

-0.06 -0.13 -0.32 -0.90

0.71 0.71 0.73 0.78

1c + 2a f 3ca 0.014 0.19 0.025 0.51 0.047 1.49 0.080 4.13

a Computed at the HF/6-31G* level. b See Scheme 13 for the numbering system. c Computed by means of eq 6. d Computed by means of eq 3. e Computed by means of eq 2. f Computed by means of eq 1.

Nu BfA [∆EAfB orb - ∆Eorb] + ∆Eorb > ∆ECoul

(4)

2

(1)

(2)

where q2 and qi are the charges of the C2 and i atoms, and R2i is the distance between C2 and the i-atom. In addition to these stabilizing interactions, there is a repulsive term which is responsible for the positive slope of the potential energy versus reaction coordinate from the reactants to the transition structure. At the initial stages of the reaction this repulsive term can be approximated as54

∆Erep ) -2( β23 S23 + β45S45)

S23

for concertedness is

(cH2 cL3 β23)2

2i

R23

+

where cHi and cLi are the expansion coefficients of HOMO and LUMO at the i-atom, respectively. Similarly, Hk and Lk are the HOMO and LUMO energies of the k-species, (A or B, see Scheme 12), respectively. The βij terms in eq 1 are the resonance integrals between the AOs of the i- and j-atoms. If we consider a stepwise reaction, the first step consists of a nucleophilic attack of azomethine A on the alkene B. In other words, the stepwise mechanism is initiated by a conjugate (Michael) addition of an enolate on an electrophilic alkene (Scheme 12). In this case, the nucleophilic addition can be described in its initial stages as the result of a combined electrostatic and orbital stabilizing interaction according to the following equation:

∆ENu orb

S45

∆Erep ∆ENu ∆Econ ∆ENu orb:∆Econ

IRC

(3)

where βij and Sij are the resonance and overlap integrals between the i- and j-atoms, respectively. Since for both the concerted and stepwise mechanisms the ∆Erep terms are similar, the nature of the mechanism is determined mainly by the alternative interactions described by eqs 1 and 2. Therefore, the condition (54) (a) Salem, L. J. Am. Chem. Soc. 1968, 90, 543. (b) Salem, L. J. Am. Chem. Soc. 1968, 90, 223.

To evaluate quantitatively eqs 1-4, several additional approximations are required. First, we have computed the resonance integrals βij by means of the Mulliken approximation:55

1 βij ≈ (β°i + β°j)Sij 2

(5)

where Sij is the corresponding overlap integral and β°i and β°j are fixed parameters for the i-and j-atoms, respectively.56 The overlap integrals can be evaluated from the analytical formulas reported by Mulliken57 for overlapping Slater-type orbitals (STOs). For σ-overlap between two 2p AOs of carbon, the analytical expression is

[

Sij ) - 1 - ξRij -

]

ξ3 R3ij ξ4 R4ij ξ2 R2ij +2 + exp(ξRij) 5 15 15

(6)

where Rij is the internuclear distance between i and j and ξ is the Slater exponent. The parameters56 used in our quantitative 2p evaluation of expressions 5 and 6 are β2p i ) βj ) - 0.28367 -1 hartree and ξ ) 1.685 bohr (See Figure S6 of the Supporting Information for additional details). The charges required for the evaluation of ∆Ecoul have been computed using the NBO charges at the different atoms of the separate reagents. Finally, the Rij values were obtained from IRC studies performed on selected reactions at the HF/6-31G* level. We have selected the 1c + 2a f 3ca and 1a + 2b f endo3ab reactions as model examples for concerted and stepwise mechanisms respectively, and the 1c + 2b f endo-3cb transformation as an example of a limiting case between both mechanisms (vide supra). The computed ∆Erep, ∆ENu, and ∆Econ energies at different points are reported in Table 7. From these data, it is clear that (55) (a) Mulliken, R. S. J. Chem. Phys. 1949, 46, 675. (b) Mulliken, R. S. J. Chem. Phys. 1952, 56, 295. (56) Dewar, M. J. S.; Zoebisch, E. G.; Healy, E. F.; Stewart, J. J. P. J. Am. Chem. Soc. 1985, 107, 3902. (57) Mulliken, R. S.; Rieke, C. A.; Orloff, D.; Orloff. H. J. Chem. Phys. 1949, 17, 1248.

6090 J. Am. Chem. Soc., Vol. 122, No. 25, 2000 in the 1c + 2a f 3ca reaction the concerted mechanism is largely preferred, since the ∆ENu terms are indeed repulsive at the first stages of the interaction. In the case of the 1c + 2b f endo-3cb reaction, the nucleophilic term becomes more significant as long as the reaction progresses. In addition, the ∆ENu orb is near 50% of the ∆Econ term, thus indicating that this is a limiting case, and that substituent and/or solvent effects can modify the mechanism from concerted to stepwise. Finally, in the case of the 1a + 2bf endo-3ab reaction the stepwise mechanism is clearly preferred, and ∆ENu orb is almost 75% of ∆Econ. In the latter cases the nitro group induces a significant polarization in the nitroalkene thus enhancing its electrophilicity. This promotes an asymmetric transfer of electron density from the azomethine to the dipolarophile. From eqs 1 and 2 we readily appreciate that if cL3 and cH2 are large and the difference of energy between the HOMO of the nitroalkene and the LUMO of the azomethine ylide increases, then ∆Econ is close to ∆ENu orb (see Table 8), thus favoring the stepwise mechanism, especially if electrostatic interactions (and eventually the electrostatic contribution of the solvation energy) are adequate. Conclusions From the combined theoretical and experimental studies reported in this paper the following conclusions can be drawn: (a) The interaction between N-metalated azomethine ylides and nitroalkenes as representative π-deficient olefins is stepwise, the first step being a conjugate nucleophilic addition in which the azomethine ylide acts as an enolate and the nitroalkene behaves as a Michael acceptor. Under favorable conditions, the corresponding zwitterionic intermediate can be trapped and its transformation into the [3 + 2] cycloadduct observed. (b) The reaction between N-unsubstituted azomethine ylides and nitroalkenes can be concerted or stepwise, depending upon the nature of the substituents. (c) The concerted or stepwise nature of the alternative reaction mechanisms can be understood and predicted using second-order perturbation theory. Since eq 4 is general for any thermal cycloaddition and approximations 5 and 6 are also quite general, the analysis can be extended to other thermal cycloadditions. Thus, within the limitations previously discussed, the nature of the mechanism can be anticipated from eqs 1 and 2, using internuclear distances of ∼3.0 Å. Experimental Section General. Commercially available compounds were used as purchased without further purification. LiClO4 (ACS reagent) was used as received (CAUTION: perchlorates are potentially explosive materials. Under the experimental conditions described below, this compound can be used safely). All solvents were dried and distilled according to standard protocols.58 Nitroalkenes 2c-g were prepared following reported procedures.59 Imines 1′d-g and 1′c were prepared as previously reported by us16 and by O’Donell,60 respectively. Cycloadducts 3dd and 3dc were characterized by comparison with the previously reported data.16 Chemical shifts in the 1H NMR and 13C NMR spectra are reported as δ values (ppm) with respect to TMS and in deuterated chloroform. Flash chromatography purifications were performed using silica gel 60-230 mesh and AcOEt/hexanes mixtures varying from 1:10 to 1:20. After the corresponding extractions, the organic layers were dried with Na2SO4 prior to evaporation of the solvent under reduced pressure. (58) Perrin, D. D.; Armarego, N. L. F. Purification of Laboratory Chemicals; Pergamon: Oxford, 1993. (59) Bourguinon, J.; LeNard, G.; Gueguiner, G. Can. J. Chem. 1985, 63, 2354. (60) O’Donell, M. J.; Polt, R. L. J. Org. Chem. 1982, 47, 2663.

ViVanco et al. General Procedures for the Reaction of the Imines 1′ and the Nitroalkenes 2. Method A. The imine 1′ (5 mmol) was dissolved in CH3CN (25 mL), and then NEt3 (0.7 mL, 5 mmol), the nitroalkene 2 (5 mmol) and LiClO4 (0.53 g, 5 mmol) were added. The progress of the reaction was monitored by TLC. After completion of the reaction, the mixture was washed with saturated aqueous NH4Cl solution (2 × 10 mL) and water (4 × 5 mL). After evaporation, the crude mixture was triturated in Et2O, and the precipitated cycloadduct was recrystallized in Et2O-hexanes. Method B. A mixture of the imine 1′ (5 mmol) and the nitroalkene 2 (5 mmol) in the appropriate solvent (25 mL) was heated at 112 °C for 24 h. After completion of the reaction, the solution was cooled and the solvent removed under reduced pressure. The different cycloadducts were separated from the reaction mixture by flash chromatography. (2S*,3R*,4S*)-5,5-Diphenyl-2-methoxycarbonyl-3-(4-methoxyphenyl)-4-nitropyrrolidine (endo-3bc): following the method A: yield 1.73 g (80%); mp 137-139 °C; IR (KBr) 3307, 1752, 1551, 1344 cm1; 1H NMR (δ ppm, CDCl3, D2O) 7.77-6.61 (m, 14H), 6.01 (d, 1H, J ) 4.0 Hz), 4.06 (dd, 1H, J ) 8.7 Hz, J′ ) 3.8 Hz), 3.96 (d, 1H, J ) 8.8 Hz), 3.73 (s, 6H); 13C NMR (δ ppm, CDCl3) 172.5, 159.8, 142.5, 129.9, 129.6, 129.3, 129.0, 128.6, 128.4, 127.4, 126.1, 114.9, 102.3, 77.9, 66.1, 58.6, 55.9, 53.2. Anal. Calcd for C25H24O5N2: C, 69.42; H, 5.60; N, 6.48. Found: C, 69.27; H, 5.63; N, 6.44. (2S*,3R*)-Methyl 2-acetamido-3-(4-methoxyphenyl)-4-nitrobutyrate (5bc). Imine 1′b (1.27 g, 5 mmol) was dissolved in CH3CN (25 mL), and then NEt3 (0.7 mL, 5 mmol), the nitroalkene 2c (0.90 g, 5 mmol), and LiClO4 (0.53 g, 5 mmol mmol) were added. After 1 h of reaction at room temperature, a 25:75 mixture of endo-3bc and imine 4bc was observed by 1H NMR. The crude reaction mixture was treated according to the general procedure A, and the resulting oily residue was dissolved in THF (15 mL) and 2 N HCl (0.8 mL). The resulting mixture was stirred for 20 min at room temperature. Usual basic workup and evaporation of the solvent gave an oily residue which was diluted with CH2Cl2 (22 mL). To the resulting solution, trifluoroacetic acid (1.5 mL, 20 mmol) was added dropwise, and the resulting mixture was stirred at room temperature for 5 h. The volatiles were evaporated, and the resulting mixture was dissolved in CH2Cl2 (8 mL). This solution was treated with NEt3 (0.42 mL, 3.12 mmol) and acetic anhydride (0.3 mL, 2.9 mmol) and stirred overnight at room temperature. The resulting mixture was washed with 1 N HCl (5 mL), NaHCO3 (saturated solution, 5 mL) and water (5 mL). The organic layer was dried, and evaporation of the solvent gave a solid which was crystallized from AcOEt/hexanes. Yield: 0.403 g (26% from 1′b); mp 146-147 °C; IR (KBr) 3308, 1742, 1660, 1554, 1252 cm1; 1H NMR (δ ppm, CDCl3) 7.10-6.82 (m, 4H), 6.15 (db, 1H, J ) 8.5 Hz), 4.87 (m, 3H), 3.86 (m, 1H), 3.78 (s, 3H), 3.57 (s, 3H), 2.03 (s, 3H); 13C NMR (δ ppm, CDCl3) 170.4, 169.9, 159.6, 129.1, 126.4, 114.4, 77.1, 55.2, 54.8, 52.5, 46.3, 23.1. Anal. Calcd for C14H18O6N2: C, 54.18; H, 5.85; N, 9.03. Found: C, 53.92; H, 5.84; N, 8.98. General Procedure for the Kinetic Measurements. In a roundbottomed flask sealed with a Liebig refrigerant the imine 1′ (5 mmol) and the nitroalkene 2 (0.5 mmol) were dissolved in the selected solvent (25 mL). The mixture was placed in an oil bath and under constant pressure. At certain time aliquots were withdrawn with a syringe, immediately quenched by cooling to 0 °C, diluted to a concentration of 10-4M, and analyzed by UV spectroscopy. Given the nature of the UV spectra, the A∞ values were calculated by means of the Kezdy-Swinbourne method.52 According to this procedure, the absorbances at t and t + τ times fulfill the following condition:

At - A ∞ ) exp(kobsτ) At+τ - A∞

(8)

At ) A∞[1 - exp(kobsτ)] + At+τ exp(kobsτ)

(9)

Therefore,

Since the first term of eq 9 is time-independent, a plot of At versus At+τ will be linear with a slope of exp(kobsτ). At the end point, At ) At+τ ) A∞. Therefore, the intersection of the line through the data points

Loss of Concertedness in Pericyclic Reactions with the 45° line gives the value of A∞. We have used τ values of ∼t1/2, as recommended,52 to provide greatest accuracy. Finally, the kobs values were determined in the usual way by means of linear regressions, as shown in Figure S4 of the Supporting Information. (2S*,3R*,4S*)-3-(4-Chlorophenyl)-2-Methoxycarbonyl-4-nitro5,5-diphenyl-pyrrolidine (endo-3bd): following the method A: yield 1.64 g (82%); mp 164-165 °C; IR (KBr) 3309, 1741, 1545, 1547, 1341 cm1; 1H NMR (δ ppm, CDCl3) 7.77-6.65 (m, 14H, arom), 6.01 (d, 1H, J ) 4 Hz), 4.08 (dd, 1H, J ) 9.0 Hz, J′ ) 4 Hz), 3.98 (d, 1H, J ) 9.1 Hz), 3.74 (s, 3H). 13C NMR (δ ppm, CDCl3) 171.5, 141.9, 141.5, 135.8, 133.8, 129.2, 129.1, 129.0, 128.6, 128.0, 127.8, 126.6, 125.3, 101.2, 77.4, 65.3, 57.6, 52.6. Anal. Calcd for C24H21O4N2Cl: C, 65.97; H, 4.85; N, 6.41. Found: C, 65.42; H, 4.89; N, 6.42. For the cyclodducts 3de, 3df, 3dg, 3dc, 3ec, 3fc, 3gc the general procedure B in toluene was followed and a mixture of compounds endoand exo- was obtained. The cycloadducts exo-3dg, endo- and exo-3fc, endo-3gc were fully characterized as their N-acyl derivatives. General Procedure for Obtaining the N-Acyl Derivatives. The corresponding cycloadduct (5 mmol) was dissolved in CH2Cl2 (50 mL), treated with TFA (50 mmol), and stirred at room temperature for 5 h. The volatiles were evaporated, and the product of the reaction dissolved in CH2Cl2 (45 mL). This solution was treated with NEt3 (7 mmol) and acetic anhydride (6.5 mmol) and stirred for 48 h at room temperature. Then it was washed with HCl 1 N (12.5 mL), saturated aqueous NaHCO3 solution (12.5 mL), and water (12.5 mL). The organic layer was dried with Na2SO4 and evaporated under reduced pressure, and the resulting syrup triturated in Et2O yielding the acetylated compound. (2S*,3R*,4S*,5S*)-2-Methoxycarbonyl-4-nitro-3,5-diphenylpyrrolidine (endo-3de): yield 0.343 g (21%); mp 120-122 °C; IR (KBr) 3314, 1735, 1545, 1370 cm1; 1H NMR (δ ppm, CDCl3) 7.44-7.25 (m, 10H), 5.27 (dd, 1H, J ) 6.5 Hz, J′ ) 3.2 Hz), 4.91 (db, 1H, J ) 6.6 Hz), 4.21 (dd, 1H, J ) 7.3 Hz, J′ ) 3.1 Hz), 4.14 (d, 1H, J ) 7.4 Hz), 3.80 (s, 3H), 3.35 (sb, 1H); 13C NMR (δ ppm, CDCl3) 171.7, 138.5, 134.4, 129.2, 128.7, 128.0, 127.4, 126.4, 96.9, 67.7, 67.3, 55.3, 52.6. Anal. Calcd for C18H18O4N2: C, 66.24; H, 5.57; N, 8.58. Found: C, 66.05; H, 5.46; N, 8.58. (2S*,3S*,4R*,5S*)-2-Methoxycarbonyl-4-nitro-3,5-diphenylpyrrolidine (exo-3de): yield 0.669 g (41%); mp 113-115 °C; IR (KBr) 3355, 1743, 1544, 1347 cm1; 1H NMR (δ ppm, CDCl3) 7.58-7.20 (m, 10H), 5.21 (t, 1H, J ) 8.0 Hz), 4.75 (db, 1H, J ) 8.2 Hz), 4.50 (d, 1H, J ) 9.0 Hz), 4.37 (t, 1H, J ) 7.7 Hz), 3.27 (s, 3H), 2.73 (sb, 1H); 13C NMR (δ ppm, CDCl3) 171.8, 137.6, 135.8, 129.0, 128.9,128.7, 128.1, 127.8, 126.8, 95.0, 67.6, 64.2, 53.7, 51.8. Anal. Calcd for C18H18O4N2: C, 66.24; H, 5.57; N, 8.58. Found: C, 65.70; H, 5.56; N, 8.58. (2S*,3R*,4S*,5S*)-2-Methoxycarbonyl-3-(4-methoxyphenyl)-4-nitro-5-phenylpyrrol-idine (endo-3df): yield 0.289 g (17%); mp 99102 °C; IR (KBr) 3317, 1738, 1545, 1373 cm1. 1H NMR (δ ppm, CDCl3) 7.38-7.14 (m, 9H), 5.24 (dd, 1H, J ) 6.4 Hz, J′ ) 3.1 Hz), 4.89 (db, 1H, J ) 6.0 Hz), 4.16 (dd, 1H, J ) 7.4 Hz, J′ ) 3.0 Hz), 4.11 (d, 1H, J ) 7.4 Hz), 3.79 (s, 3H), 3.34 (sb, 1H), 2.35 (s, 3H); 13C NMR (δ ppm, CDCl3) 171.7, 137.8, 135.5, 134.4, 129.9, 128.6, 127.3, 126.4, 97.0, 67.6, 67.3, 55.0, 52.5, 21.0. Anal. Calcd for C19H20O4N2: C, 67.03; H, 5.93; N, 8.23. Found: C, 66.80; H, 5.90; N, 8.15. (2S*,3S*,4R*,5S*)-2-Methoxycarbonyl-3-(4-methylphenyl)-4-nitro-5-phenylpyrrol-idine (exo-3df): yield 0.714 g (42%); mp 130132 °C; IR (KBr) 3367, 1738, 1549, 1365 cm1; 1H NMR (δ ppm, CDCl3) 7.59-7.10 (m, 9H), 5.18 (t, 1H, J ) 8.0 Hz), 4.74 (d, 1H, J ) 8.3 Hz), 4.47 (d, 1H, J ) 9.0 Hz), 4.33 (t, 1H, J ) 8.3 Hz), 3.30 (s, 3H), 2.71 (sb, 1H), 2.29 (s, 3H); 13C NMR (δ ppm, CDCl3) 172.0, 137.9, 137.7, 132.7, 129.4, 129.0, 128.9, 127.7, 126.9, 95.2, 67.6, 64.2, 53.5, 51.8, 21.0. Anal. Calcd for C19H20O4N2: C, 67.03; H, 5.93; N, 8.23. Found: C, 66.91; H, 5.85; N, 8.41. (2S*,3R*,4S*,5S*)-3-Trifluoromethylphenyl-2-methoxycarbonyl4-nitro-5-phenylpyrrolidine (endo-3dg): yield 0.394 g (20%); mp 155-157 °C; IR (KBr) 3353, 1736, 1555, 1321 cm1; 1H NMR (δ ppm, CDCl3) 7.69-7.25 (m, 9H), 5.27 (dd, 1H, J ) 6.7 Hz, J′ ) 4.0 Hz), 4.92 (tb, 1H), 4.29 (dd, 1H, J ) 7.6 Hz, J′ ) 3.9 Hz), 4.13 (tb, 1H), 3.81 (s, 3H), 3.30 (sb, 1H); 13C NMR (δ ppm, CDCl3) 172.0, 143.0, 134.9, 129.7, 129.5, 128.8, 127.2, 127.0, 126.9, 97.2, 68.4, 67.9, 55.3,

J. Am. Chem. Soc., Vol. 122, No. 25, 2000 6091 53.5. Anal. Calcd for C19H17O4N2F3: C, 57.86; H, 4.35; N, 7.10. Found: C, 57.55; H, 4.39; N, 7.10. (2S*,3S*,4R*,5S*)-1-Acetyl-3-trifluoromethylphenyl-2-methoxycarbonyl-4-nitro-5-phenylpyrrolidine (acetyl derivative of exo3dg): overall yield from 1′d and 2g 0.196 g (9%); mp 200-202 °C; IR (KBr) 1745, 1660, 1551, 1327 cm1; 1H NMR (δ ppm, CDCl3) 7.697.37 (m, 9H), 5.70 (dd, 1H, J ) 11.9 Hz, J′ ) 8.4 Hz), 5.36 (d, 1H, J ) 8.5 Hz), 5.16 (d, 1H, J ) 9.21 Hz), 4.43 (t, 1H, J ) 10.7 Hz), 3.36 (s, 3H), 1.68 (s, 3H); 13C NMR (δ ppm, CDCl3) 170.7, 170.6, 137.6, 135.1, 129.7,129.6, 128.3, 126.9, 126.0, 125.9, 125.8, 92.9, 66.7, 63.6, 52.2, 49,1, 22.7. Anal. Calcd for C21H19O5N2F3: C, 57.80; H, 4.40; N, 6.42. Found: C, 57.60; H, 4.33; N, 6.42. (2S*,3R*,4S*,5S*)-2-Methoxycarbonyl-3-(4-methoxyphenyl)-5-(4methyl-phenyl)-4-nitropyrrolidine (endo-3ec): yield 0.370 g (20%); mp 96-98 °C; IR (KBr) 3297, 1736, 1546, 1356 cm1; 1H NMR (δ ppm, CDCl3) 7.25-6.87 (m, 8H), 5.20 (dd, 1H, J ) 6.6 Hz, J′ ) 3.3 Hz), 4.85 (db, 1H, J ) 6.1 Hz), 4.14 (dd, 1H, J ) 7.3 Hz, J′ ) 3.3 Hz), 4.08 (d, 1H, J ) 7.4 Hz), 3.81 (s, 3H), 3.79 (s, 3H), 3.30 (sb, 1H), 2.32 (s, 3H); 13C NMR (δ ppm, CDCl3) 171.9, 159.3, 138.5, 131.4, 130.6, 129.4, 128.6, 126.3, 114.6, 97.3, 67.6, 67.5, 55.3, 54.9, 52.6, 21.2. Anal. Calcd for C20H22O5N2: C, 64.84; H, 6.00; N, 7.56. Found: C, 64.45; H, 6.00; N, 7.62. (2S*,3S*,4R*,5S*)-2-Methoxycarbonyl-3-(4-methoxyphenyl)-5-(4methyl-phenyl)-4-nitropyrrolidine (exo-3ec): yield 0.426 g (23%); mp 141-144 °C; IR (KBr) 3290, 1722, 1546, 1335 cm1; 1H NMR (δ ppm, CDCl3) 7.45-6.80 (m, 8H), 5.14 (t, 1H, J ) 8.1 Hz), 4.68 (tb, 1H, J ) 8.1 Hz), 4.43 (dd, 1H, J ) 9.1 Hz, J′ ) 5.4 Hz), 4.31 (t, 1H, J ) 8.4 Hz), 3.76 (s, 3H), 3.32 (s, 3H), 2.67 (tb, 1H), 2.36 (s, 3H); 13C NMR (δ ppm, CDCl3) 172.0, 159.2, 138.6, 134.6, 129.6, 128.9, 127.6, 126.6, 114.0, 95.2, 67.3, 64.1, 55.1, 53.0, 51.8, 21.1. Anal. Calcd for C20H22O5N2: C, 64.84; H, 6.00; N, 7.56. Found: C, 64.54; H, 6.04; N, 7.56. (2S*,3R*,4S*,5S*)-1-Acetyl-5-(4-chlorophenyl)-2-methoxycarbonyl-3-(4-methoxyphenyl)-4-nitropyrrolidine (acetyl derivative of endo-3fc): overall yield from 1′f and 2c 0.173 g (8%); mp 220 °C; IR (KBr) 1747, 1657, 1551, 1398 cm1; 1H NMR (δ ppm, CDCl3) 7.686.83 (m, 8H), 5.56 (dd, 1H, J ) 11.9 Hz, J′ ) 8.5 Hz), 5.31 (d, 1H, J ) 8.7 Hz), 5.07 (d, 1H, J ) 9.4 Hz), 4.29 (dd, 1H, J ) 11.9 Hz, J′ ) 9.5 Hz), 3.77 (s, 3H), 3.40 (s, 3H), 1.67 (s, 3H); 13C NMR (δ ppm, CDCl3) 171.8, 171.2, 160.6, 137.3, 130.5, 129.5, 129.0, 128.5, 123.0, 115.1, 94.2, 66.7, 64.5, 55.9, 52.9, 49.8, 23.4. Anal. Calcd for C21H21O6N2Cl: C, 58.26; H, 4.90; N, 6.47. Found: C, 57.94; H, 4.94; N, 6.45. (2S*,3S*,4R*,5S*)-1-Acetyl-5-(4-chlorophenyl)-2-methoxycarbonyl-3-(4-methoxy-phenyl)-4-nitropyrrolidine (acetyl derivative of exo-3fc): overall yield from 1′f and 2c 0.184 g (8.5%); mp 214-215 °C; IR (KBr) 1742, 1661, 1556, 1393 cm1; 1H NMR (δ ppm, CDCl3) 7.68-6.83 (m, 8H), 5.56 (tb, 1H, J ) 8.6 Hz), 5.31 (d, 1H, J ) 8.4 Hz), 5.07 (d, 1H, J ) 9.4 Hz), 4.29 (tb, 1H, J ) 10.1 Hz), 3.78 (s, 3H), 3.40 (s, 3H), 1.67 (s, 3H); 13C NMR (δ ppm, CDCl3) 171.9, 171.2, 160.6, 137.3, 130.5, 129.5,129.0, 128.5, 123.0, 115.1, 94.2, 66.7, 64.5, 55.9, 52.9, 49.8, 23.4. Anal. Calcd for C21H21O6N2Cl: C, 58.26; H, 4.90; N, 6.47. Found: C, 57.95; H, 4.92; N, 6.38. (2S*,3R*,4S*,5S*)-1-Acetyl-5-(4-trifluoromethylphenyl)-2-methoxycarbonyl-3-(4-methoxyphenyl)-4-nitropyrrolidine (acetyl derivative of endo-3gc): overall yield from 1′g and 2c 0.222 g (9.5%); mp 200-204 °C; IR (KBr) 1742, 1661, 1554, 1326 cm1; 1H NMR (δ ppm, CDCl3) 7.98-6.84 (m, 8H), 5.55 (s, 1H), 5.12 (d, 1H, J ) 10.5 Hz), 5.04 (d, 1H, J ) 6.1 Hz), 3.95 (d, 1H, J ) 6.2 Hz), 3.91 (d, 1H, J ) 6.4 Hz), 3.78 (s, 3H), 3.72 (s, 3H), 1.92 (s, 3H); 13C NMR (δ ppm, CDCl3) 171.8, 170.4, 160.0, 141.5, 128.8, 127.2, 126.7, 126.5, 121.8, 114.5, 95.9, 65.7, 62.2, 55.2, 52.8, 48.5, 21.9. Anal. Calcd for C22H21O6N2F3: C, 56.65; H, 4.55; N, 6.00. Found: C, 56.50; H, 4.55; N, 6.05.

Acknowledgment. This work has been supported by the Gobierno Vasco-Eusko Jaurlaritza (Projects EX-1998-126 and PI-1998-116 and Fondo de Cooperacio´n Aquitania/Euskadi/ Navarra) and by the Provincial Government of Gipuzkoa (Gipuzkoako Foru Aldundia). We also thank Professor Carlos

6092 J. Am. Chem. Soc., Vol. 122, No. 25, 2000 Ubide (UPV/EHU) for helpful suggestions on the kinetic measurements. Supporting Information Available: Tables of absolute energies and ZPVEs of all the stationary points discussed in the text and numerical data used in the elaboration of Figures

ViVanco et al. S1 and S2, Figures showing the evolution of the bond distances against IRC and variation of β against R (PDF). Crystallographic data for endo-3bc and 5bc (CIF). This material is available free of charge via the Internet at http://pubs.acs.org. JA9945360